Cooperative Ordering of Collagen Triple Helices in the Dense State

Apr 19, 2007 - Aurélien Tidu , Djida Ghoubay-Benallaoua , Claire Teulon , Sophie Asnacios , Kate Grieve , François Portier , Marie-Claire Schanne-Kl...
2 downloads 0 Views 299KB Size
Langmuir 2007, 23, 6411-6417

6411

Cooperative Ordering of Collagen Triple Helices in the Dense State F. Gobeaux,†,‡ E. Belamie,*,† G. Mosser,† P. Davidson,‡ P. Panine,§ and M.-M. Giraud-Guille† Chimie de la Matie` re Condense´ e, UMR 7574 CNRSsUniVersite´ Pierre et Marie Curie, ENSCPsEcole Pratique des Hautes Etudes, 12 rue CuVier, 75005 Paris, Laboratoire de Physique des Solides, UMR 8502, CNRS, UniVersite´ Paris Sud, Baˆ t. 510, 91405 Orsay Cedex, and High Brilliance Beamline ID2, European Synchrotron Radiation Facility, BP 220, F-38043 Grenoble Cedex, France ReceiVed January 12, 2007. In Final Form: March 8, 2007 Extracellular matrixes such as bone, skin, cornea, and tendon have ordered structures comprised for the most part of collagen, an elongated protein of well-defined dimensions and composition. Here we show how the cooperative ordering of collagen triple helices in the dense fluid state is exploited to produce dense ordered collagen matrixes. The spontaneous formation of a birefringent phase occurs at critical concentrations that increase from 50-60 to 80-85 mg/mL as the acetic acid concentration of the solvent increases from 5 to 500 mM. We studied by small-angle X-ray scattering (SAXS) the local liquidlike positional order across the isotropic/anisotropic phase transition by unwinding the cholesteric phase with moderate shearing stress. Interparticle scattering gives rise to a broad interference peak. The average distance between triple helices, dav, is thus estimated and decreases linearly as a function of φ-1/2 from 12.7 ( 0.9 nm (22.5 mg/mL) to 5.0 ( 0.6 nm (166.4 mg/mL). Equilibrium concentrations and the order parameter of the nematic phase agree reasonably well with theoretical predictions for semiflexible macromolecules. Striated fibrils with a high degree of alignment were obtained by fine-tuning the delicately balanced electrostatic interactions, which yielded strong elastic gels with a hierarchical organization very similar to that of major biological tissues. Typical Bragg reflections corresponding to the 67 nm period characteristic of collagen fibrils in biological tissues were recorded by SAXS with ordered collagen matrixes reconstituted in vitro.

Introduction Connective tissues are complex biological systems with integrated properties and functions and represent a good source of inspiration for material scientists.1 The biosynthesis of type I collagen involves a controlled spatiotemporal sequence of protein expression, folding, transport, and secretion into the extracellular space. A unique feature of collagen is the highly periodic nature of its assembly from the amino acid sequence to its long-range organization in soft and hard tissues. Collagen molecules are formed by three polypeptide chains folded together into a triple helical structure approximately 300 nm in length and 1.5 nm in diameter.2 After procollagen has been secreted by cells, the nonhelical propeptides present at both ends of the molecule are removed by specific enzymes. The resulting mature collagen is insoluble at physiological pH and precipitates into fibrils, known to consist of periodic three-dimensional assemblies of triple helices with a long-range axial periodicity. The molecules are assembled axially head-to-tail and set laterally 1.2 nm apart on a quasihexagonal lattice3,4 with a 67 nm stagger of the centers of mass.5 This regularly staggered structure induces periodic variations of electron density visible in electron microscopy as cross-striations and by X-ray diffraction as sharp Bragg peaks. Finally, at the micrometer to millimeter scale most of the tissues mentioned above exhibit a very regular, often periodic, suprafibrillar organization. An example is the lamellar morphology of compact * Corresponding author. E-mail: [email protected]. Tel.: +33 (0) 4 67 16 34 44. Fax: +33 (0) 4 67 16 34 70. † CNRS-EPHE-UPMC. ‡ CNRS-Universite ´ Paris Sud. § European Synchrotron Radiation Facility. (1) Sanchez, C.; Arribart, H.; Giraud-Guille, M.-M. Nat. Mater. 2005, 4, 277288. (2) Woodhead-Galloway, J. Studies in Biology; E. Arnold: London, 1980; Vol. 117. (3) Hulmes, D. J. S.; Miller, A. Nature 1979, 282, 878-880. (4) Hulmes, D. J. S. J. Struct. Biol. 2002, 137, 2-10. (5) Kadler, K. E.; Holmes, D. F.; Trotter, J. A.; Chapman, J. A. Biochem. J. 1996, 316, 1-11.

bone, where the collagen-apatite composite exhibits a longrange helical geometry,6 which helps to limit the propagation of cracks.7 The structural analogy between major extracellular matrixes and mesophases8,9 suggests that a transient liquid-crystal state could direct the formation of highly ordered biological tissues. This hypothesis has been formulated for a number of biological materials, and lyotropic liquid-crystalline states have been described in vivo.10-16 Cooperative ordering in suspensions of elongated macromolecules or particles can be addressed theoretically by considering that the predominant interaction is the excluded volume between impenetrable particles.17 We show here, through in vitro experiments, that self-assembly processes may account for major aspects of collagen organization in connective tissues.18 On the basis of theoretical models of a lyotropic mesophase of charged semiflexible spherocylinders with a large aspect ratio (L/D),19,20-21 we show that the intrinsic properties of collagen triple helices can yield the complex ordered structures observed in biological tissues. In particular, the helical (6) (a) Giraud-Guille, M.-M. Curr. Opin. Solid State Mater. Sci. 1998, 3, 221-227. (b) Giraud-Guille, M.-M. Calcif. Tissue Int. 1988, 42, 167-180. (7) Peterlik, H.; Roschger, P.; Klaushoffer, K.; Fratzl, P. Nat. Mater. 2006, 5, 52-55. (8) Bouligand, Y. Tissue Cell 1972, 4, 189-217. (9) Neville, A. C. Biology of fibrous composites, deVelopment beyond the cell membrane; Cambridge University Press: Cambridge, U.K., 1993. (10) Neville, A. C. Mol. Cryst. Liq. Cryst. 1981, 76, 279-286. (11) Neville, A. C.; Luke, B. M. J. Cell Sci. 1971, 8, 93-109. (12) Knight, D. P.; Feng, D.; Stewart, M.; King, E. Philos. Trans. R. Soc. London, B 1993, 341, 419-436. (13) Kerkam, K.; Viney, C.; Kaplan, D.; Lombardi, S. Nature 1991, 349, 596-598. (14) Vollrath, F.; Knight, D. P. Nature 2001, 410, 541-548. (15) Livolant, F.; Bouligand, Y. Chromosoma 1978, 68 21-44 (16) Willison, J. H. M.; Abeysekera, R. M. J. Polym. Sci. 1988, 26, 71-75. (17) Onsager, L. Ann. N.Y. Acad. Sci. 1949, 51, 627-659. (18) Belamie, E.; Gobeaux, F.; Mosser, G.; Giraud-Guille, M.-M. J. Phys.: Condens. Matter. 2006, 18, 115-129. (19) Khokhlov, A. R.; Semenov, A. N. Physica A 1982, 112, 605-614. (20) Odijk, Th. Macromolecules 1986, 19, 2313-2329. (21) Chen, Z. Y. Macromolecules 1993, 26, 3419-3423.

10.1021/la070093z CCC: $37.00 © 2007 American Chemical Society Published on Web 04/19/2007

6412 Langmuir, Vol. 23, No. 11, 2007

organization of collagen in bone6 is reminiscent of the chiral structure of cholesteric liquid crystals. The microscopic textures and the localization of the phase transition confirm22-24 that highly concentrated collagen solutions exhibit properties similar to those of polymer liquid crystals. The microscopic structure of the nematic phase is described by probing the local order with small-angle X-ray scattering (SAXS). Finally, we show how a sol-gel transition can easily be triggered to stabilize the liquidcrystal phase into a dense ordered fibrillar material that mimics the hierarchical organization of biological tissues with interesting properties. Experimental Section Collagen Preparation. Type I collagen was extracted from Wistar rat tails. The tendons were excised in a sterile cell culture hood and washed with PBS (phosphate-buffered saline) to remove cells and traces of blood. The tissues were then soaked in a 4 M NaCl solution to lyse the remaining cells and precipitate some of the high-molecularweight proteins. After being rinsed again with PBS, the clean tendons were solubilized in 500 mM acetic acid. The solution was clarified by a first centrifugation step at 41000g for 2 h. Proteins other than type I collagen were selectively precipitated in 300 mM NaCl and removed by centrifugation at 41000g for 3 h. Collagen was retrieved from the supernatant by precipitation in 600 mM NaCl and centrifugation at 3000g for 45 min. The pellets were solubilized in the final solvent (acetic acid, 5, 50, or 500 mM) and thoroughly dialyzed against the same solvent to completely remove NaCl. The solutions were kept at 4 °C and centrifuged at 41000g for 4 h before use. The collagen concentration was determined by assessing the hydroxyproline amount25 and purity assessed by SDS-PAGE electrophoresis. Given the quantities necessary for our experiments, we could not afford too drastic a purification, and traces of collagen, types III and V, and proteoglycans can still be present. Sample Preparation and Texture Analysis. The solvent, 5, 50, or 500 mM acetic acid, was slowly evaporated from a large volume of dilute collagen solution (5 mg/mL) in sterile conditions. In the concentration range near the transition, samples were taken at regular intervals, tested for their birefringence, and titrated. The optical birefringence and textures of the collagen solutions were investigated to determine the critical concentrations and the large-scale molecular organization of the samples. For this purpose, samples of the collagen solutions were sucked into 200 µm thick flat capillary tubes and observed between crossed polarizers on a Nikon Eclipse E600Pol microscope, equipped with a DXM 1200 CCD camera to capture high-resolution pictures. Dilute collagen solutions are isotropic and appear dark when viewed between crossed polarizers. In contrast, concentrated collagen solutions exhibit birefringent textures in polarized light microscopy. We chose solvent evaporation to prepare concentrated samples for mainly two reasons: (i) We wanted the results to be comparable with previous studies.23,24 (ii) The method, although time-consuming, is much faster than dialysis to prepare a large number of samples in a concentration series. The major drawback of evaporation is that the solvent composition changes because water evaporates faster than acetic acid. To take this effect into account, dilute acetic acid solutions of concentrations 5, 50, and 500 mM were evaporated in the same conditions as the collagen samples and titrated. The acetic acid concentrations determined in this way were used to build up the phase diagram in Figure 2. It must be noted that the overall trend of the graph is not modified by this correction. Experiments to identify textures were also performed with thin glass chambers continuously fed with a dilute collagen solution at one end. Slow evaporation of the solvent at the other end induces the controlled increase in volume fraction. The solutions first develop (22) Murthy, N. S. Biopolymers 1984, 23, 1261-1267. (23) Giraud-Guille, M.-M. J. Mol. Biol. 1992, 224, 861-873. (24) Mosser, G.; Anglo, A.; Helary, C.; Bouligand, Y.; Giraud-Guille, M. M. Matrix Biol. 2006, 25, 3-13. (25) Bergaman, I; Loxley, R. Anal. Chem. 1963, 35, 1961-1963.

Gobeaux et al. a strongly birefringent layer at the air/liquid interface. As evaporation proceeds, the anisotropic domain widens and a new interface is created between an anisotropic, optically birefringent, fluid and an isotropic liquid phase. A detailed description of the setup is given in ref 24. The interface progression speed is on the order of 10 µm/h, and since cholesteric globules are found as far as 50 µm from the isotropic/anisotropic interface, we estimate that the characteristic lifetime of the transitory equilibrium is several hours. Fibrillogenesis. Concentrated liquid-crystalline samples, in either capillaries or glass chambers, were placed in the presence of ammonia vapors at room temperature.26,27 Gaseous ammonia slowly dissolves in water and neutralizes the acetic acid, causing the pH to rise locally, without diluting the samples. White, dense opaque gels were obtained after 2-24 h depending on the samples’ geometries. After the gels were rinsed in deionized water or PBS until neutral pH was reached, the resulting hydrated elastic solids were directly submitted to X-ray synchrotron radiation or prepared for TEM observations. SAXS Characterization. SAXS experiments were performed on the ID02 station at the European Synchrotron Radiation Facility (ESRF). We investigated the scattering from collagen solutions in a wide q range, from 0.1 to 10 nm-1, using monochromatic synchrotron radiation (λ ) 0.0995 nm) and a sample-to-detector distance of either 1 or 3 m. Two-dimensional SAXS patterns were recorded using an X-ray image intensifier (Thomson) lens coupled to a fast read-out CCD (FreLoN camera) with a typical exposure time of 1 s. Appropriate corrections were carried out,28 and the data were analyzed with Fit2d (ESRF) and Image (Laboratoire de Physique des Solides (LPS), Universite´ Paris Sud) softwares. Concentrated strongly birefringent samples were placed in the gap of a polycarbonate cylindrical shear cell29 and subjected to moderate controlled shear stress to produce nematic monodomains. This classic alignment procedure has been applied to many other systems and is comparable to the application of an external electric or magnetic field.30 The samples used were obtained by slow evaporation of an initial dilute solution in 500 mM acetic acid. The azimuthal profiles were taken at the q value corresponding to the intensity maximum (qmax) for a given sample, and the intensity was averaged over 15 CCD pixels. After fibrillogenesis with ammonia vapors, the typical 67 nm periodicity of fibrillar type I collagen was detected with the dense gels inserted in X-ray cylindrical glass capillary tubes and placed directly in the synchrotron beam. TEM Observations. Dense collagen matrixes and tendons extracted from rat tail were fixed at 4 °C successively in 2.5% glutaraldehyde and in 2% osmium tetraoxide, both for 60 min in a cacodylate buffer at pH 7.4. The samples were then progressively dehydrated through successive ethanol baths (50%, 70%, 95%, 100%) before being embedded in Araldite for ultrathin sectioning. The 60 nm thick sections were deposited on copper grids, stained with phosphotungstic acid, and observed with a Philips CM12 electron microscope operated at 120 kV.

Results Textures and Sequence of Phases. Highly concentrated collagen solutions were examined in polarized light to assess the first-order nature of the isotropic/cholesteric (I/N*) transition. In concentration gradients, birefringent droplets exhibiting a striped pattern typical of a cholesteric phase were observed to coexist with an isotropic phase (Figure 1a). At even higher volume fractions, the birefringent droplets merge to yield a homogeneous fingerprint pattern (Figure 1b). The succession of bright and dark stripes arises from the helical organization of collagen (26) Besseau, L. A; Giraud-Guille, M.-M. J. Mol. Biol. 1995, 251, 197-202. (27) Giraud-Guille, M.-M.; Besseau, L. Connect. Tissue Res. 1997, 37, 183193. (28) Narayanan, T.; Diat, O.; Bosecke, P. Nucl. Instrum. Methods Phys. Res., Sect. A 2001, 467, 1005-1009. (29) Panine, P.; Gradezielski, M.; Narayanan, T. ReV. Sci. Instrum. 2003, 74, 2451-2455. (30) Lemaire, B. J.; Davidson, P.; Ferre´, J.; Jamet, J. P.; Panine, P.; Dozov, Y.; Jolivet, J. P. Phys. ReV. Lett. 2002, 88, 125507.

CooperatiVe Ordering of Collagen Triple Helices

Langmuir, Vol. 23, No. 11, 2007 6413

Figure 1. Typical liquid-crystalline textures observed in polarized light microscopy. (a) Birefringent globules with the striped pattern characteristic of the cholesteric phase emerge from the isotropic medium. The coexistence of the two phases indicates a first-order transition. Bar 20 µm. (b) At higher concentrations, the homogeneous anisotropic phase exhibits the typical fingerprint pattern. Bar 10 µm. (c, d) The anisotropic phase is sucked into a capillary tube. The sample appears dark when the nematic director is oriented along the direction of either polarizer (0°) and bright otherwise with a maximum at 45°. Bar 100 µm. (e) Banded textures were observed with highly concentrated collagen solutions sucked into flat capillary tubes. Bar 100 µm. The polarizer and analyzer orientations are indicated in (c).

molecules in the chiral nematic phase, also called cholesteric or N*. The resulting fingerprint pattern with characteristic defects is typical of the textures obtained with other cholesteric liquid crystals, both lyotropic and thermotropic. The cholesteric pitch is typically a few micrometers. Birefringent samples can be strongly aligned by sucking them into capillary tubes (Figures 1c,d). The moderate shear stress (on the order of 10 s-1) generated by the flow is large enough to unwind the chiral texture and produce homogeneous uniaxial single domains. This also happens when shearing stress is applied in a cylindrical shear cell (see the SAXS experiments below). Highly concentrated viscous samples sucked into capillary tubes display banded textures, with wide bright domains separated by thin extinction lines (Figure 1e). Inserting a retardation plate in the optical path produces colors that indicate undulations of the nematic director in the plane of observation. Samples were prepared in acetic acid with increasing collagen concentrations to locate the phase transition boundaries. Isotropic samples close to the I/N* transition, when sucked into capillary tubes, exhibit a strong flow birefringence that disappears in a matter of seconds as the sample relaxes and the transient alignment is lost. At higher concentrations, the solutions remain birefringent after shearing. We did not observe the bulk phase separation of the isotropic and cholesteric phases in sheared samples. We can thus locate the dilute boundary of the coexistence domain, i.e., the concentration, Ci, at which samples start to exhibit birefringence at rest. The critical concentrations increase from 5060 to 80-85 mg/mL as the acetic acid concentration increases from 5 to 500 mM (Figure 2). Small-Angle X-ray Scattering. Strongly anisotropic SAXS patterns were obtained with highly concentrated collagen solutions submitted to very moderate shear (Figure 3a). The diffuse scattering intensity is higher in the vertical direction, indicating that the elongated collagen molecules are oriented horizontally along the flow direction. The shear-induced alignment does not relax when shearing is stopped, and the strong uniaxial orientation of the sample remains, in good agreement with optical observation of birefringence. In the anisotropic phase, collagen molecules are aligned along an average direction, n (nematic director), with an orientational distribution function (ODF), f(a), a being

Figure 2. Sequence of phases. Collagen concentration series solutions were prepared in acetic acid at different concentrations. Isotropic samples (9) appear dark or exhibit a transient birefringence. Beyond the critical concentration, the cholesteric samples appear strongly birefringent (0). The thick line is a guide to the eye and indicates the region of the I/N* transition.

the unit vector associated with each individual molecule.31 The first nonzero moment of the ODF, which quantifies the strength of the ordering, is called the nematic order parameter, S. Taking a section through the 2D pattern at a fixed q modulus gives a profile of the intensity as a function of the azimuthal angle ψ (Figure 3b), the shape of which is directly related to the molecules’ orientational distribution. To estimate S, the profiles were fitted according to a procedure described elsewhere.32 A flat profile corresponds to a disordered sample (S ) 0), whereas a sharp peak indicates strong alignment of the molecules (S f 1). Typically, for a 90 mg/mL collagen solution in 500 mM acetic acid, optically birefringent at rest but close to the phase boundary, the order parameter S is estimated to be 0.45 ( 0.1. Interparticle scattering gives rise to a broad interference peak due to liquidlike order, whose maximum is located at a q value, qmax, inversely (31) Maier, W.; Saupe, A. Z. Naturforsch. 1958, 13, 564-566. (32) Davidson, P.; Petermann, D.; Levelut, A. M. J. Phys. II 1995, 5, 113131.

6414 Langmuir, Vol. 23, No. 11, 2007

Gobeaux et al.

Figure 3. Small-angle X-ray scattering of concentrated collagen solutions spanning the I/N* transition: (a) SAXS pattern of a nematic collagen solution (91 mg/mL) recorded under moderate shear (γ˘ ) 8 s-1), (b) experimental azimuthal profile taken along the curved dotted line at qmax as indicated in (a) (symbols) and fitted according to ref 32 (line), (c) linear profiles I ) f(q) taken along the vertical direction of the SAXS pattern at concentrations spanning the whole range investigated.

Figure 4. Small-angle X-ray scattering of an isotropic collagen solution (37 mg/mL). (a) Relaxed state. The SAXS pattern (γ˘ ) 0 s-1) is isotropic. (b) The same solution under shearing (γ˘ ) 400 s-1) yields a strongly anisotropic pattern indicating shear-induced order. (c) Azimuthal profiles of a 37 mg/mL solution under increasing shear stress (0, 1, 16, 64, and 300 s-1). The profile is flat at γ˘ ) 0 and becomes sharper as the shear stress is increased. (d) Values of S estimated by fitting the intensity azimuthal profiles32 as a function of γ˘ .

proportional to the average distance, dav, between the triple helices (qmax ) 2π/dav). Radial profiles of the intensity scattered at small angles, I ) f(q), display maxima (Figure 3c) and were used in combination with the 2D patterns to determine the dav values plotted in Figure 5. SAXS patterns of dilute samples are isotropic at rest but turn anisotropic under shear (Figure 4a,b). In agreement with optical observations of flow birefringence, the samples appear to relax very fast, as the scattering patterns quickly revert to isotropic once the shearing stress is stopped. Shear-induced ordering of the isotropic phase is revealed by the azimuthal profiles becoming sharper as the shear stress is increased (Figure 4c). The order parameter measured under shear rapidly increases from 0 to 0.1 for a very moderate shear stress of 1 s-1. Isotropic samples also exhibit an interference peak, revealing local positional order. The values of dav determined with nematic and isotropic samples are shown together in Figure 5. The average distance, dav, decreases with increasing volume fraction, φ, following a straight line that goes through the origin, as a function of φ-1/2. The relatively large width of the peaks in Figure 3c indicates that the local positional order only extends to nearest neighbors. The correlation length, ξ, of the liquidlike order, calculated as ξ ) 2π/∆q (with ∆q the peak’s fwhm), does not exceed 25 nm

Figure 5. Dilution law of collagen in acetic acid. Left axis: average distance between collagen molecules (dav ) 2π/qmax) as a function of φ-1/2 (cholesteric, b; isotropic, O). Right axis: corresponding correlation length of the local order, ξ (*). The labels indicate concentrations in milligrams per milliliter.

(at 22.5 mg/mL) and tends to decrease as the triple helices become more tightly packed, down to 8.5 nm at 91 mg/mL (dav ) 6.45 nm). Biomimetic Fibrillar Matrixes. We subjected highly concentrated collagen solutions to ammonia vapor to induce

CooperatiVe Ordering of Collagen Triple Helices

Langmuir, Vol. 23, No. 11, 2007 6415

Figure 6. Typical 67 nm period of collagen fibrils in dense matrixes: TEM observations of rat tail tendon (a) and an 18 wt % dense aligned collagen matrix (b). The SAXS pattern in (c) corresponds to a dense gel reconstituted in vitro as in (b).

precipitation into fibrils to form a strong gel.26,27 Gaseous NH3 that dissolves in the aqueous phase raises the pH by neutralizing the acetic acid, while keeping the collagen volume fraction unchanged. For comparison, ultrathin sections of rat tendon observed by TEM displayed in Figure 6a show bundles of crossstriated fibrils. The same pattern is actually observed with sections of a dense gel (18 wt %) prepared by reassembly in vitro (Figure 6b). The SAXS patterns recorded with the dense gels formed in capillary tubes are slightly anisotropic at the smallest angles, indicating the presence of large aligned aggregates. The wideangle interference peak, corresponding to lateral interactions between triple helices in the fibrils, gives an intermolecular distance of about 1.5 nm (preliminary data, not shown). Sharp Bragg rings corresponding to harmonics of the 67 nm periodicity of the fibrils also show a small but significant reinforcement in the vertical direction. The azimuthal spread measured on a radial intensity profile is typically less than 40°.

Discussion Study of the Textures in Polarized Light. The fingerprint patterns observed with slowly evaporated collagen solutions are typical of a cholesteric phase and were presented in previous papers.23,24 The coexistence of cholesteric globules with an isotropic phase shown in Figure 1a is a clear indication that the transition is of the first-order type. Transitory thermodynamic equilibrium is thus reached despite the concentration gradient induced by evaporation. Controlled evaporation in thin glass chambers has proven a very suitable method to produce and observe biphasic samples in a reproducible manner. This experimental setup is based on the continuous injection of collagen in a confined space. Although concentration is obtained by slow evaporation of the solvent at one interface, the system is continuously enriched with a fresh dilute solution. Therefore, it is likely that diffusion of acetic acid allows equilibration throughout the chamber. The banded pattern presented in Figure 1e looks like those reported during evaporation of collagen solutions either between the slide and coverslip or with an open air interface27,33,34 and with DNA fragments.35 Banded textures are ubiquitous in polymer liquid crystals subjected to shear36,37 and have been shown to arise from periodic undulations of the director around the shear direction38 due to stress relaxation. There is strong evidence in the literature supporting the idea of a liquid-crystalline state preceding the formation of many biological materials. In some instances, the mesophase was identified in vivo, and the presence (33) Maeda, H. Langmuir 1999, 15, 8505-8513. (34) Maeda, H. Langmuir 2000, 16, 9977-9982. (35) Livolant, F. J. Phys. (Paris) 1987, 48, 1051-1066. (36) Yan, N. X.; Labes M. M. Macromolecules 1994, 27, 7843-7845. (37) Wang, J.; Labes, M. M. Macromolecules 1992, 25, 5790-5793. (38) Han, W. H.; Rey, A. D. Macromolecules 1995, 28, 8401-8405.

of distortions induced by shear during biological processing was demonstrated in particular for the collagenous material of the dogfish egg case12,39 and the spider’s silk protein.13,14 Interestingly, in tendon, the collagen fibrils follow an undulating organization called crimp40-43 that exhibits some similarities with the instabilities observed with polymer liquid crystals under external fields.36,44 Whether or not the crimp in tendon might originate from similar mechanisms as those inducing in vitro the formation of banded texture is not yet demonstrated. However, shear-induced banding of collagen solutions could be exploited to prepare materials with specific properties inspired from the crimp structure of tendon. Isotropic/Cholesteric Phase Transition. According to the phase diagram presented in Figure 2, the transition from a disordered isotropic phase to a chiral nematic one occurs at critical concentrations that increase from 50-60 to 80-85 mg/mL as the initial acetic acid concentration of the solvent increases from 5 to 500 mM. The original theory of the isotropic/nematic (I/N) transition was developed by Onsager for rigid rods.17 Khokhlov and Semenov19 proposed an extension that accounts for the particles’ flexibility, characterized by the ratio R ) L/P, where L is the contour length and P the persistence length. The authors showed that flexibility would increase the critical concentrations and lower the order parameter. On the basis of the marked flexibility of collagen triple helices, Chen’s simulations21 for long semiflexible molecules were used to estimate the concentration, Ci, expected at the transition. The repulsive potential between the like-charged macromolecules extends over a few Debye lengths and was treated as a scaling factor by considering an effective electrostatic diameter, Deff, instead of the bare geometrical one, Dbare.19-21,45-48 The net charge of collagen was estimated on the basis of its amino acid sequence (Swiss-Prot protein database). Taking into account a hydroxylation level of lysines of about 20%,49 we found an excess of typically 232 positive charges in 5 mM acetic acid (pH 3.52), or 257 in 500 mM acetic acid (pH 2.53). In these conditions and for a total collagen concentration of 100 mg/mL, for instance, the corresponding values of Deff are respectively 3.96 and 3.73 nm, more (39) Knight, D. P.; Hu, X. W.; Gathercole, L. J.; Rusaoue¨n-Innocent, M.; Ho, M.-W.; Newton, R. Philos. Trans. R. Soc. London, B 1996, 351, 1205-1222. (40) Diamant, J.; Keller, A.; Baer, E.; Litt, M.; Arridge, R. G. C. Proc. R. Soc. London, Ser. B 1972, 180, 293-315. (41) Gathercole, L. J.; Keller, A. Micron 1978, 9, 83-89. (42) Vidal, B. Micron 2003, 34, 423-432. (43) Feitosa, V. L. C.; Reis, F. P.; Esquisatto, M. A. M; Joazeiro, P. P.; Vidal, B. C.; Pimentel, E. R. Micron 2006, 37, 518-525. (44) Curtis, R. F. Macromolecules 1986, 19, 2431-2435. (45) Stroobants, A.; Lekkerkerker, H. N. W.; Odijk, Th. Macromolecules 1986, 19, 2232-2238. (46) Lee, S. D. J. Chem. Phys. 1987, 87, 4972-4974. (47) Tang, J.; Fraden, S. Liq. Cryst. 1995, 19, 459-467. (48) Sato, T.; Teramoto, A. Physica A 1991, 176, 72-86. (49) Uzawa, K.; Yeowell, H. N.; Yamamotoa, K.; Mochidaa, Y.; Tanzawab, H.; Yamauchi, M. Biochem. Biophys. Res. Commun. 2003, 305, 484-487.

6416 Langmuir, Vol. 23, No. 11, 2007

than 2Dbare. Using published values of 160 nm,50-52 112 nm,51 and 57 nm53 for the persistence length, we obtained critical concentrations, Ci, of, respectively, 15.4, 28.9, and 111.8 mg/ mL for samples of the 500 mM acetic acid series. This illustrates the high sensitivity of the model to flexibility,19,21 and the predicted concentrations essentially bracket the experimental data of 8085 mg/mL. The agreement is nevertheless significantly closer than the value of 2.95 mg/mL predicted by Onsager’s model for rigid rods in the same conditions. As seen in Figure 2, the experimental coexistence concentrations tend to increase as the acetic acid concentration is increased. This trend is also well predicted by the theory as Ci increases from about 98.6 mg/mL in 5 mM acetic acid to 111.8 mg/mL in 500 mM acetic acid, with P ) 57 nm. This is true for every persistence length and is a direct consequence of the decrease in Deff being essentially due to the higher ionic strength at higher acetic acid concentration. More accurate quantitative testing of the models would require the determination of P in our experimental conditions and in particular at very high collagen concentration. In the description of the I/N transition in statistical physics models, two entropic terms contribute to the total free energy and vary inversely as the degree of order increases. The entropy of mixing favors the random orientation of the rods and is highest in the isotropic phase. The excluded volume entropy decreases the total free energy as the average angle between neighboring particles becomes smaller, in favor of tighter packing in the nematic phase. A typical feature of the I/N first-order transition is the discontinuity in the order parameter, S. For perfectly rigid rods, S is predicted to be 0 in the isotropic phase, 0.8 in the nematic phase at coexistence, and 1 for an ideally aligned sample. The value of 0.45 ( 0.1 determined from the SAXS patterns of collagen aligned samples is very close to that expected for semiflexible particles. The predicted value21 of S varies from 0.477 with the highest value of the persistence length (P ) 160 nm) to 0.492 if the particles are considered more flexible (P ) 57 nm). Therefore, theories for elongated semiflexible particles with an effective diameter controlled by surface charges describe and predict reasonably well the isotropic/cholesteric phase transition of collagen solutions at high volume fractions. Local Order in the Dense Solutions. For a nematic phase of very anisotropic charged rodlike particles, the dependence of dav on the volume fraction, also called the “swelling law”, is of the form54-61 d ) kφ-1/2. This is expected for solutions, both anisotropic and isotropic, well above the overlap concentration, c* ) 1/L3. If we further assume that the molecules’ axes are set locally on a 2D hexagonal lattice, the slope (k ) 0.825 D) gives a triple helix diameter of 1.9 nm, in reasonable agreement with the value of 1.5 nm given in the literature. The correlation length, (50) Claire, K; Pecora, R. J. Phys. Chem. 1997, 101, 746-753. (51) Saito, T.; Iso, N.; Mizuno, H.; Onda, N.; Yamato, H.; Odashima, H. Biopolymers 1982, 21, 715-728. (52) Henry, F.; Nestler, M.; Hvidt, S.; Ferry, J. D.; Veis, A. Biopolymers 1983, 22, 1747-1758. (53) Hofmann, H.; Voss, T.; Ku¨hn, K.; Engel, J. J. Mol. Biol. 1984, 172, 325-343. (54) Groot, L. C. A.; Kuil, M. E.; Leyte, J. C.; Van der Maarel, J. R. C.; Heenan, R. K.; King, S. M.; Jannick, G. Liq. Cryst. 1994, 17, 263-276. (55) Maier, E. E.; Krause, R.; Deggelmann, M.; Hagenbu¨chle, M.; Weber, R.; Fraden, S. Macromolecules 1992, 25, 1125-1133. (56) Purdy, K. R.; Dogic, Z.; Fraden, S.; Ru¨hm, A.; Lurio, L.; Mochrie, S. G. J. Phys. ReV. E 2003, 67, 31708. (57) Belamie, E.; Davidson, P.; Giraud-Guille, M. M. J. Phys. Chem. B 2004, 108, 14991-15000. (58) De Gennes, P. G.; Pincus, P.; Velasco, R. M.; Brochard, F. J. Phys. (Paris) 1976, 37, 1461-1473. (59) Keates, P.; Mitchell, G. R.; Peuvrel, E. Polymer 1992, 33, 3298-3301. (60) Borsali, R.; Rinaudo, M.; Noirez, L. Macromolecules 1995, 28, 10851088. (61) Hagenbu¨chle, M.; Weyerich, B.; Deggelmann, M.; Graf, C.; Krause, R.; Maier, E. E.; Schultz, S. F.; Klein, R.; Weber, R. Physica A 1990, 169, 29-41.

Gobeaux et al.

ξ, inferred from the linear profiles is at most 25 nm at a concentration where dav ) 12.7 nm, which clearly shows that the order extends only to nearest neighbors. Moreover, the interference peak broadens as the concentration is increased and even vanishes at 166 mg/mL, suggesting that the electrostatic repulsions predominant at lower volume fraction may be overcome by attractive interactions at high packing densities. Although counterintuitive, such a vanishing of the correlation peak in suspensions of rodlike molecules stabilized by electrostatic repulsion is actually rather the rule than the exception.57,62-64 Alignment of the Isotropic Phase under Shear. Expectedly, the isotropic phase shows no birefringence at rest (Figure 1a), and the corresponding SAXS patterns show no long-range orientational order (Figure 4a). However, when submitted to even moderate shear stress, the solutions exhibit transient birefringence that disappears as soon as the shearing is stopped. The order parameters inferred from the corresponding scattering patterns, taken during the application of shear, show a sharp increase of S at moderate values of shear stress and then a slow increase, theoretically tending toward 1. Such a strong susceptibility to shear probably reveals a tendency to local orientational ordering in the isotropic phase, as previously reported for fdvirus61 and TMV55 suspensions at concentrations above c*. S ) 0.35 is the highest value that could be reached in the shear stress range accessible experimentally (600 s-1) (Figure 4d). This is therefore the practical limit of alignment for anisotropic materials prepared from isotropic collagen solutions at c < 80 mg/mL in 500 mM acetic acid. Biomimetic Fibrillar Matrixes Exhibit the Typical 67 nm X-ray Diffraction Signal. Negative staining of the samples reveals in TEM the succession of gap and overlap zones (Figure 6a,b) along the collagen fibrils present in the dense gels as in rat tail tendons. The regular striations arise from variations in the electron density along the fibrils, and their periodicity (D ) 67 nm) is due to the regularly staggered arrangement of adjacent macromolecules.2,65 Because of the extreme regularity along the fibril axis and because of lateral liquidlike order, an analogy has been made with lamellar liquid crystals.66,67 We were able to identify Bragg reflections from the 67 nm period (Figure 6c) in dense anisotropic gels reconstituted in vitro. Collagen fibrils in vivo usually have a finite lateral size and are locally packed closely parallel to each other (Figure 6a). TEM observations show a similar local alignment of fibrils in our samples, confirmed by SAXS, with an azimuthal spread of the Bragg reflections of about 40°. It should be noted that, in the absence of shear, the gels exhibit long-range helicity because of the intrinsic chiral nature of the fluid phase.24,26 This long-range chirality preserved in the solid may confer interesting mechanical properties to reassembled materials.7 Unlike the orientational order in the nematic phase dominated by cylindrically averaged repulsions in the nanometer range, the formation of ordered fibrils involves details of the amino acid distribution, on the angstrom scale. Repulsive interactions arising from the excess of positive charges are maximal at acidic pH. Between pH 6 and pH 9 the net charge is only slightly positive and the isoelectric point is reached around pH 9.3. In the neutral(62) Wang, L.; Bloomfield, V. A. Macromolecules 1991, 24, 5791-5795. (63) Orts, W. J.; Godbout, L.; Marchessault, R. H.; Revol, F.-F. Macromolecules 1998, 31, 5717-5725. (64) Pelletier, O.; Davidson, P.; Bourgaux, C.; Livage, J. Europhys. Lett. 1999, 48, 53-59. (65) Bruns, R. R.; Trelstad, R. L.; Gross, J. Science 1973, 181, 269-271. (66) Hukins, D. W. L.; Woodhead-Galloway, J. Mol. Cryst. Liq. Cryst. 1977, 41, 33-39. (67) Fratzl, P.; Fratzl-Zelman, N.; Klaushofer, K. Biophys. J. 1993, 64, 260266.

CooperatiVe Ordering of Collagen Triple Helices

to-basic pH range, not only are striated fibrils formed but the concentrated collagen solutions turn into strong elastic gels. It is likely that the fibrils act as reticulation nodes and ensure the strong coherence of the solid. We have noticed that the dense gels do not readily redissolve in acetic acid, possibly indicating chemical reticulation, as suggested, among others, by Suarez et al., who studied the reversibility of fibrillogenesis by synchrotron radiation X-ray scattering.68 Reconstituted fibrils are often obtained from dilute solutions69 or, in contrast, after thorough dehydration.70,71 In the present work, combined TEM and SAXS studies prove that the delicate molecular interaction between tightly packed triple helices required for the formation of striated fibrils is reached in highly concentrated collagen solutions, despite their highly viscoelastic behavior. X-ray scattering has been used previously to study the structure of in vitro reconstituted fibrils and its dependence on pH, osmotic pressure, temperature, etc. We now have the opportunity to study how typical fibrils’ X-ray signals (∼1.5 and 67 nm) appear in dense liquid-crystalline collagen solutions. Abundant literature has been devoted to the formation of striated fibrils in vitro and in vivo (see, for instance, refs 4, 5, 69, and 71-75). However, questions remain as to how fibrils are formed in vivo, how their diameter is regulated, and how their composition is adapted to meet the functional requirements of different tissues, although a number of additional macromolecules, collagenous or not, have already been shown to play specific roles.73-76 The large-scale organization of collagen, undulated in tendon,40-43 orthogonal in cornea,77 and helical in bone,6 is even more difficult to explain. Besides the control exerted by cells74,75,78 on collagen secretion, simple physicochemical parameters can largely influence the 3D ordering of the protein in the extracellular space. A better understanding of the influence of factors such as concentration, ionic strength, and pH, or the effect of external fields such as shearing, might help answer questions regarding the in vivo processes of collagen assembly.

Conclusions The outstanding solution properties of collagen make it a versatile system for tailoring ordered materials with hierarchical (68) Suarez, G.; Oronsky, A. L.; Bordas, J.; Koch, M. H. J. Proc. Natl. Acad. Sci. U.S.A. 1985, 82, 4693-4696. (69) Wood, G. C.; Keech, M. K. Biochem. J. 1960, 75, 588-598. (70) Leikin, S.; Rau, D. C.; Parsegian, V. A. Proc. Natl. Acad. Sci. U.S.A. 1994, 91, 276-280. (71) Kuznetsova, N.; Leikin, S. J. Biol. Chem. 1999, 274, 36083-36088. (72) Comper, W. D.; Veis, A. Biopolymers 1977, 16, 2113-2131. (73) Gelman, R. A.; Williams, B. R.; Piez, K. A. J. Biol. Chem. 1979, 254, 180-186. (74) Canty, E. G.; Kadler, K. E. Comp. Biochem. Physiol., Part A 2002, 133, 979-985. (75) Canty, E. G.; Kadler, K. E. J. Cell Sci. 2005, 118, 1341-1353. (76) Holmes, D. F.; Watson, R. B.; Chapman, J. A.; Kadler, K. E. J. Mol. Biol. 1996, 261, 93-97. (77) Trelstad, R. L.; Coulombre, A. J. J. Cell Biol. 1971, 50, 840-858. (78) Canty, E. G.; Lu, Y.; Meadows, R. S.; Shaw, M. K.; Holmes, D. F.; Kadler, K. E. J. Cell Biol. 2004, 165, 553-563.

Langmuir, Vol. 23, No. 11, 2007 6417

levels of organization. We were able to assess the first-order nature of the isotropic/cholesteric phase transition in highly concentrated collagen solutions. Mapping out the phase diagram revealed that the critical concentrations increase by 50%, from 50-60 to 80-85 mg/mL, as the acetic acid concentration of the solvent increases from 5 to 500 mM. In the latter case, the theoretical predictions range from 15.4 to 111.8 mg/mL depending on the value considered for the persistence length of collagen, reasonably bracketing the experimental data. Variations of the average distance between triple helices deduced from SAXS patterns, dav, as a function of the volume fraction indicate local 2D hexagonal packing of the triple helices. The diffuse scattering shows only liquidlike order characterized by a low value of the correlation length. In fact, the highest value taken by ξ is 25 nm when dav ) 12.7 nm, and therefore, lateral order is limited to nearest neighbors. However, concentrated anisotropic solutions exhibit a relatively high nematic order parameter of 0.45, as expected from modeling. The triple helices thus have the intrinsic ability to strongly align locally while remaining several nanometers apart, in the liquid-crystal state. Orientational ordering of soluble collagen in vitro occurs spontaneously without requiring any protein activity or cellular action. Obviously the situation in vivo is far more complex, where chaperone activity, plasma membrane folding, and the secreting cells’ cytoskeletons can play a role in the extracellular trafficking and processing of collagen. In vitro, increasing the pH of concentrated liquidcrystalline solutions results in the formation of strong gels comprised of locally aligned cross-striated fibrils. Typical Bragg reflections, corresponding to the 67 nm period characteristic of collagen fibrils in biological tissues, were recorded by SAXS with ordered collagen matrixes reconstituted in vitro. Because we know that the fluid state follows self-assembly principles described by statistical physics, we can apply theoretical models to fine-tune the local and long-range organization of collagen in solution. It is therefore possible to prepare a wide variety of ordered hierarchical materials, mimicking biological connective tissues, by applying the same processes. However, apart from fibrillar collagens, living tissues are also comprised of other important macromolecules, specific cell types, and in some cases a mineral phase, all participating in their structures, specificities, and functionalities. These aspects are currently under investigation, in particular the behavior of specialized cells such as fibroblasts79 or osteoblasts in the presence of reconstituted fibrillar matrixes. Acknowledgment. We thank J. Doucet and D. Durand for very helpful discussions. We also thank Rachel Carol for reading and revising the manuscript. LA070093Z (79) Helary, C.; Ovtracht, L.; Coulomb, B.; Godeau, G.; Giraud-Guille, M.-M. Biomaterials 2006, 27, 4443-4452.